首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 218 毫秒
1.
The trinuclear platinum compound [{trans‐PtCl(NH3)2}2(μ‐trans‐Pt(NH3)2{NH2(CH2)6NH2}2)]4+ (BBR3464) belongs to the polynuclear class of platinum‐based anticancer agents. These agents form in DNA long‐range (Pt,Pt) interstrand cross‐links, whose role in the antitumor effects of BBR3464 predominates. Our results show for the first time that the interstrand cross‐links formed by BBR3464 between two guanine bases in opposite strands separated by two base pairs (1,4‐interstrand cross‐links) exist as two distinct conformers, which are not interconvertible, not only if these cross‐links are formed in the 5′‐5′, but also in the less‐usual 3′‐3’ direction. Analysis of the conformers by differential scanning calorimetry, chemical probes of DNA conformation, and minor groove binder Hoechst 33258 demonstrate that each of the four conformers affects DNA in a distinctly different way and adopts a different conformation. The results also support the thesis that the molecule of antitumor BBR3464 when forming DNA interstrand cross‐links may adopt different global structures, including different configurations of the linker chain of BBR3464 in the minor groove of DNA. Our findings suggest that the multiple DNA interstrand cross‐links available to BBR3464 may all contribute substantially to its cytotoxicity.  相似文献   

2.
《中国化学会会志》2017,64(8):934-939
This study evaluated the structural, electronic and thermochemical properties of an anticancer active molecule, i.e. trans‐bis‐(3‐aminoflavone)dichloridoplatinum(II) (trans‐Pt(3‐af)2Cl2; TCAP) in the gas and solution phases. The polarizable continuum model (PCM) model was used to perform the required calculations in five solvents with different polarities. Moreover, the dependencies of energetic aspects, structural, thermodynamic parameters and frontier orbital energies of the complex were also examined. Dependencies of the frequency shifts of u(CO), u(NH) and 195Pt Chemical shifts on the solvent dielectric were investigated by Kirkwood–Bauer–Magat equation (KBM). The energies of platinum d‐orbitals and formal electron configurations of Pt atom were calculated by natural bond analysis (NBO).  相似文献   

3.
The first and second substitution reactions binding of the anticancer drug trans‐[Pt((CH3)2C?NOH)((CH3)2CHNH2)Cl2] to purine bases were studied computationally using a combination of density functional theory and isoelectric focusing polarized continuum model approach. Our calculations demonstrate that the trans monoaqua and diaqua reactant complexes (RCs) can generate either trans‐ or cis‐monoadducts via identical or very similar trans trigonal‐bipyramidal transition‐state structures. Furthermore, these monoadducts can subsequently close by coordination to the adjacent purine bases to form 1,2‐intrastrand Pt‐DNA adducts and eventually distort DNA in the same way as cisplatin. Thus, it is likely that the transplatin analogues have the same mechanism of anticancer activity as cisplatin. For the first substitutions, the activation free energies of monoaqua complexes are always lower than that of diaqua complexes. The lowest activation energy for monoaqua substitutions is 16.2 kcal/mol for guanine and 16.5 kcal/mol for adenine, whereas the lowest activation energy for diaqua substitutions is 17.1 kcal/mol for guanine and 25.9 kcal/mol for adenine. For the second substitutions, the lowest activation energy from trans‐monoadduct to trans‐diadduct is 19.1 kcal/mol for GG adduct and 20.7 kcal/mol for GA adduct, whereas the lowest activation energy from cis‐monoadduct to cis‐diadduct is 18.9 kcal/mol for GG adduct and 18.5 kcal/mol for GA adduct. In addition, the first and second substitutions prefer guanine over adenine, which is explained by the remarkable larger complexation energy for the initial RC in combination with lower activation energy for the guanine substitution. Overall, the hydrogen‐bonds play an important role in stabilizing these species of the first and second substitutions. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

4.
Some platinum boryl complexes of the type trans‐[(Cy3P)2Pt(Cl){B(Cl)R}] ( 1 : R = NMe2, 2 : R = Mes, 3 : R = tBu) were synthesized by oxidative addition of the corresponding dichloroboranes to [Pt(PCy3)2]. All the compounds were characterized by multinuclear NMR spectroscopy in solution. Furthermore, a single crystal analysis was acquired from 2 , that confirms the strong trans‐influence of this boryl ligand.  相似文献   

5.
Platinum diam(m)ine complexes, such as cisplatin, are successful anticancer drugs, but suffer from problems of resistance and side‐effects. Photoactivatable PtIV prodrugs offer the potential of targeted drug release and new mechanisms of action. We report the synthesis, X‐ray crystallographic and spectroscopic properties of photoactivatable diazido complexes trans,trans,trans‐[Pt(N3)2(OH)2(MA)(Py)] ( 1 ; MA=methylamine, Py=pyridine) and trans,trans,trans‐[Pt(N3)2(OH)2(MA)(Tz)] ( 2 ; Tz=thiazole), and interpret their photophysical properties by TD‐DFT modelling. The orientation of the azido groups is highly dependent on H bonding and crystal packing, as shown by polymorphs 1 p and 1 q . Complexes 1 and 2 are stable in the dark towards hydrolysis and glutathione reduction, but undergo rapid photoreduction with UVA or blue light with minimal amine photodissociation. They are over an order of magnitude more potent towards HaCaT keratinocytes, A2780 ovarian, and OE19 oesophageal carcinoma cells than cisplatin and show particular potency towards cisplatin‐resistant human ovarian cancer cells (A2780cis). Analysis of binding to calf‐thymus (CT), plasmids, oligonucleotide DNA and individual nucleotides reveals that photoactivated 1 and 2 form both mono‐ and bifunctional DNA lesions, with preference for G and C, similar to transplatin, but with significantly larger unwinding angles and a higher percentage of interstrand cross‐links, with evidence for DNA strand cross‐linking further supported by a comet assay. DNA lesions of 1 and 2 on a 50 bp duplex were not recognised by HMGB1 protein, in contrast to cisplatin‐type lesions. The photo‐induced platination reactions of DNA by 1 and 2 show similarities with the products of the dark reactions of the PtII compounds trans‐[PtCl2(MA)(Py)] ( 5 ) and trans‐[PtCl2(MA)(Tz)] ( 6 ). Following photoactivation, complex 2 reacted most rapidly with CT DNA, followed by 1 , whereas the dark reactions of 5 and 6 with DNA were comparatively slow. Complexes 1 and 2 can therefore give rapid potent photocytotoxicity and novel DNA lesions in cancer cells, with no activity in the absence of irradiation.  相似文献   

6.
A novel trans‐platinum(II) complex bearing one dimethylamine (dma) and one methylamine (ma) ligand, namely trans‐[PtCl2(dma)(ma)], recently synthesised and characterised in our laboratory, displayed relevant antiproliferative properties in vitro, being more active than the parent complex, trans‐[PtCl2(dma)(ipa)], which has isopropylamine (ipa) in place of methylamine. We have analysed comparatively the solution behaviour of these two complexes under various experimental conditions, and investigated their reactivity with horse heart cytochrome c by mass spectrometry, inductively coupled plasma–optical emission spectroscopy (ICP‐OES), 2D [1H,15N],[1H,13C] HSQC and [1H,1H] NOESY NMR. Some important changes that occurred in the [1H,13C] HSQC NMR spectrum of cytochrome c treated with trans‐[PtCl2(dma)(ma)] in water, after two days’ incubation, most probably arose from direct platinum coordination to the protein side chain; this was proved conclusively by [1H,1H] NOESY NMR and [1H,15N] HSQC NMR measurements. Met65 was identified as the primary Pt binding site on cytochrome c. Electrospray mass spectrometry (ESIMS) results provided evidence for extensive platinum–protein adduct formation. A fragment of the [Pt(amine)(amine′)] type was established to be primarily responsible for protein metalation. ICP‐OES analysis revealed that these trans‐platinum(II) complexes bind preferentially to the serum proteins albumin and transferrin rather than to calf thymus DNA. Pt binding to DNA was found to be far lower than in the case of cisplatin. The implications of the results for the mechanism of action of novel cytotoxic trans‐platinum complexes are discussed.  相似文献   

7.
Both cis‐ and trans‐di­chloro­bis­(di­phenyl ­sulfide)­platinum(II), [PtCl2(C12H10S)2], crystallize as mononuclear pseudo‐square‐planar complexes. In the cis compound, the Pt—Cl distances are 2.295 (2) and 2.319 (2) Å, and the Pt—S distances are 2.280 (2) and 2.283 (2) Å. In the trans compound, Pt is located on a centre of inversion and the Pt—Cl and Pt—S distances are 2.2786 (15) and 2.3002 (12) Å, respectively.  相似文献   

8.
The polar phosphanyl‐carboxamide, 1′‐(diphenylphosphanyl)‐1‐[N‐(2‐hydroxyethyl)carbamoyl]ferrocene ( 1 ), reacts readily with hydrogen peroxide and elemental sulfur to give the corresponding phosphane‐oxide and phosphane‐sulfide, respectively, and with platinum(II) and palladium(II) precursors to afford various bis(phosphane) complexes [MCl2( 1 ‐κP)2] (M = trans‐Pd, trans‐Pt and cis‐Pt). The anticancer activity of the compounds was evaluated in vitro with the complexes showing moderate cytotoxicities towards human ovarian cancer cells. Moreover, the biological activity was found to be strongly influenced by the stereochemistry, with trans‐[PtCl2( 1 ‐κP)2] being an order of magnitude more active than the corresponding cis isomer. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

9.
A stable trans‐(alkyl)(boryl) platinum complex trans‐[Pt(BCat′)Me(PCy3)2] (Cat′=Cat‐4‐tBu; Cy=cyclohexyl=C6H11) was synthesised by salt metathesis reaction of trans‐[Pt(BCat′)Br(PCy3)2] with LiMe and was fully characterised. Investigation of the reactivity of the title compound showed complete reductive elimination of Cat′BMe at 80 °C within four weeks. This process may be accelerated by the addition of a variety of alkynes, thereby leading to the formation of the corresponding η2‐alkyne platinum complexes, of which [Pt(η2‐MeCCMe)(PCy3)2] was characterised by X‐ray crystallography. Conversion of the trans‐configured title compound to a cis derivative remained unsuccessful due to an instantaneous reductive elimination process during the reaction with chelating phosphines. Treatment of trans‐[Pt(BCat′)Me(PCy3)2] with Cat2B2 led to the formation of CatBMe and Cat′BMe. In the course of further investigations into this reaction, indications for two indistinguishable reaction mechanisms were found: 1) associative formation of a six‐coordinate platinum centre prior to reductive elimination and 2) σ‐bond metathesis of B? B and C? Pt bonds. Mechanism 1 provides a straightforward explanation for the formation of both methylboranes. Scrambling of diboranes(4) Cat2B2 and Cat′2B2 in the presence of [Pt(PCy3)2], fully reductive elimination of CatBMe or Cat′BMe from trans‐[Pt(BCat′)Me(PCy3)2] in the presence of sub‐stoichiometric amounts of Cat2B2, and evidence for the reversibility of the oxidative addition of Cat2B2 to [Pt(PCy3)2] all support mechanism 2, which consists of sequential equilibria reactions. Furthermore, the solid‐state molecular structure of cis‐[Pt(BCat)2(PCy3)2] and cis‐[Pt(BCat′)2(PCy3)2] were investigated. The remarkably short B? B separations in both bis(boryl) complexes suggest that the two boryl ligands in each case are more loosely bound to the PtII centre than in related bis(boryl) species.  相似文献   

10.
Transition‐metal complexes bearing fluorinated phosphane and thiolate ligands has been an area of study in recent years and the chemical context of the current work is related to the metal‐assisted functionalization of fluorinated derivatives. The cis and trans isomers of the square‐planar complex bis[(pentafluorophenyl)diphenylphosphane‐κP]bis(2,3,5,6‐tetrafluorobenzenethiolato‐κS)platinum(II), [Pt(C6HF4S)2{P(C6H5)2(C6F5)}2], have been crystallized from a single chromatographic fraction and characterized by X‐ray diffraction analysis. The stabilization of the cis isomer results from weak intramolecular π‐stacking interactions and possibly from the formation of a C—F…Pt contact, characterized by an F…Pt separation of 2.957 (6) Å. The natural bond orbital analysis (NBO) for this isomer confirms that the corresponding F → Pt charge transfer accounts for 6.92 kcal mol−1 in the isomer stabilization. Such interactions are not present in the centrosymmetric trans isomer.  相似文献   

11.
Several trans‐platinum(II) complexes, of the type R′? {Pt(PBu3)2}? R″? {Pt(PR3)2}? R′, where R′ and R″ are groups derived from a series of aromatic alkynes and diynes, have been prepared and characterized. Extensive spectroscopic data for these and other known related complexes are presented. A more precise structural study of trans‐Pt(C≡CC6H4C≡CPh)2(PBu3)2 (cf. Z. Kristallogr. 1998; 213: 483) is reported. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

12.
A series of azine‐bridged dinuclear platinum(II) complexes of the type [{trans‐Pt(NH3)2(OH2)}2(μ‐azn)](ClO4)4 (where azn = pyrazine (pzn, Pt1 ), 2,3‐dimethylpyrazine (2,3‐pzn, Pt2 ), and 2,5‐dimethylpyrazine (2,5‐pzn, Pt3 )) were synthesized to investigate the influence of the bridging azine ligand on the reactivity of the platinum(II) centers. The pKa values of the complexes were determined via acid–base titration, and the rate of substitution of the aqua moiety by a series of neutral nucleophiles, viz. thiourea (TU), 1,3‐dimethyl‐2‐thiourea (DMTU), and 1,1,3,3‐tetramethyl‐2‐thiourea (TMTU), was determined under pseudo‐first‐order conditions as a function of concentration and temperature using standard spectrophotometric techniques. The introduction of the methyl groups to the bridging azine linker in Pt2 and Pt3 leads to a moderate increase in the pKa values obtained for the first and second deprotonation steps, respectively, as a result of the increased σ‐donor capacity of the bridging azine ligand trans to the aqua moiety. A comparison of the rate constants, k1 and k2, at 298 K, obtained for the substitution of the aqua moieties from Pt1 , Pt2 , and Pt3 by TU, shows that the introduction of the σ‐donating methyl groups on the bridging azine ligand in Pt2 and Pt3 results in a corresponding decrease in the reactivity, by ca. five times for the first substitution step and ca. 10 times for the second substitution step. Density functional theory calculations at the B3LYP/LACVP** level of theory for the complexes demonstrate that the introduction of electron‐donating methyl groups results in (i) increased steric hindrance over the metal centers and (ii) decreased the positive charge on the metal center and increases energy separation of the frontier molecular orbitals (EHOMOELUMO) of the ground‐state platinum(II) complexes, leading to a less‐reactive metal center. © 2011 Wiley Peiodicals, Inc. Int J Chem Kinet 43: 161–174, 2011  相似文献   

13.
The platina‐β‐diketone [Pt2{(COMe)2H}2(µ‐Cl)2] ( 1 ) was found to react with monodentate phosphines to yield acetyl(chloro)platinum(II) complexes trans‐[Pt(COMe)Cl(PR3)2] (PR3 = PPh3, 2a ; P(4‐FC6H4)3, 2b ; PMePh2, 2c ; PMe2Ph, 2d ; P(n‐Bu)3, 2e ; P(o‐tol)3, 2f ; P(m‐tol)3, 2g ; P(p‐tol)3, 2h ). In the reaction with P(o‐tol)3 the methyl(carbonyl)platinum(II) complex [Pt(Me)Cl(CO){P(o‐tol)3}] ( 3a ) was found to be an intermediate. On the other hand, treating 1 with P(C6F5)3 led to the formation of [Pt(Me)Cl(CO){P(C6F5)3}] ( 3b ), even in excess of the phosphine. Phosphine ligands with a lower donor capability in complexes 2 and the arsine ligand in trans‐[Pt(COMe)Cl(AsPh3)2] ( 2i ) proved to be subject to substitution by stronger donating phosphine ligands, thus forming complexes trans‐[Pt(COMe)Cl(L)L′] (L/L′ = AsPh3/PPh3, 4a ; PPh3/P(n‐Bu)3, 4b ) and cis‐[Pt(COMe)Cl(dppe)] ( 4c ). Furthermore, in boiling benzene, complexes 2a – 2c and 2i underwent decarbonylation yielding quantitatively methyl(chloro)platinum(II) complexes trans‐[Pt(Me)Cl(L)2] (L = PPh3, 5a ; P(4‐FC6H4)3, 5b ; PMePh2, 5c ; AsPh3, 5d ). The identities of all complexes were confirmed by 1H, 13C and 31P NMR spectroscopy. Single‐crystal X‐ray diffraction analyses of 2a ·2CHCl3, 2f and 5b showed that the platinum atom is square‐planar coordinated by two phosphine ligands (PPh3, 2a ; P(o‐tol)3, 2f ; P(4F‐C6H4)3, 5b ) in mutual trans position as well as by an acetyl ligand ( 2a, 2f ) and a methyl ligand ( 5b ), respectively, trans to a chloro ligand. Single‐crystal X‐ray diffraction analysis of 3b exhibited a square‐planar platinum complex with the two π‐acceptor ligands CO and P(C6F5)3 in mutual cis position (configuration index: SP‐4‐3). Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

14.
Summary The halogen bridges of the dimeric, cyclometallated trimesityl-arsine and -phosphine complexes of palladium(II) and platinum(II), where M=Pd or Pt and E=P or As have been replaced with pyrazolate groups to give the corresponding and less symmetric pyrazolato-bridged complexes, where M=Pd or Pt, E=P or As, Pz=pyrazolato anion, and M=Pd, E=As, Pz=3,5-dimethylpyrazolato anion. In the case of the palladium complexes,1H. n.m.r. clearly indicates the presence of only one isomer which is most likely to have thetrans configuration while the platinum complexes are mixtures of bothcis andtrans forms.Part VI, ref. 3c  相似文献   

15.
The crystal structure of the title compound, trans‐[PtI2(C6H12N3P)2], describes one of the few platinum(II) complexes containing two of the water‐soluble 1,3,5‐tri­aza‐7‐phosphaadamantane ligands reported to date. The complex crystallizes on an inversion centre with the most important bond lengths and angles being Pt—P 2.3128 (12) Å, Pt—I 2.6022 (6) Å, P—Pt—I 90.94 (3)° and P′—Pt—I 89.06 (3)°.  相似文献   

16.
Eight platinum acetylide complexes have been synthesized and characterized. The catalytic properties of these complexes in curing silicone rubber by hydrosilylation have been tested. Among the complexes tested, trans‐Pt(PPh3)2[―C≡CC(CH3)2OSi(CH3)3] 2 (2), trans‐Pt(PPh3)2[―C≡CC(CH3)2OSi(CH2CH3)3]2 (3), trans‐Pt(PPh3)2[―C≡CC(CH3)2OSiPh(CH3)2]2 (4), trans‐Pt(PPh3)2[―C≡C(C6H10)OSi(CH3)3]2 (6), trans‐Pt(PPh3)2[―C≡C(C6H10)OSi(CH2CH3)3]2 (7), and trans‐Pt(PPh3)2[―C≡C(C6H10)OSiPh(CH3)2]2 (8) exhibited sufficiently long pot‐lives (15 days) at room temperature and short silicone rubber curing times of 10–35 min at 100°C or 1–5 min at 120°C. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

17.
1H, 13C and 15N NMR studies of gold(III), palladium(II) and platinum(II) chloride complexes with picolines, [Au(PIC)Cl3], trans‐[Pd(PIC)2Cl2], trans/cis‐[Pt(PIC)2Cl2] and [Pt(PIC)4]Cl2, were performed. After complexation, the 1H and 13C signals were shifted to higher frequency, whereas the 15N ones to lower (by ca 80–110 ppm), with respect to the free ligands. The 15N shielding phenomenon was enhanced in the series [Au(PIC)Cl3] < trans‐[Pd(PIC)2Cl2] < cis‐[Pt(PIC)2Cl2] < trans‐[Pt(PIC)2Cl2]; it increased following the Pd(II) → Pt(II) replacement, but decreased upon the transcis‐transition. Experimental 1H, 13C and 15N NMR chemical shifts were compared to those quantum‐chemically calculated by B3LYP/LanL2DZ + 6‐31G**//B3LYP/LanL2DZ + 6‐31G*. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

18.
A series of dinuclear cycloplatinated(II) complexes with general closed formula of [Pt2Me2(C^N)2(μ‐P^P)] (C^N = 2‐vinylpyridine (Vpy), 2,2′‐bipyridine N‐oxide (O‐bpy), 2‐(2,4‐difluorophenyl)pyridine (dfppy); P^P = 1,1‐bis(diphenylphosphino)methane (dppm), N,N‐bis(diphenylphosphino)amine (dppa)) are reported. The complexes were characterized by means of NMR spectroscopy. Due to the presence of dppm and dppa with short backbones as bridging ligands, two platinum centres are located in front of each other in these complexes so a Pt…Pt interaction is established. Because of this Pt…Pt interaction, the complexes have bright orange colour under ambient light and are able to strongly emit red light under UV light exposure. These strong red emissions originate from a 3MMLCT (metal–metal‐to‐ligand charge transfer) electronic transition. In most of these complexes, the emissions have unstructured bell‐shaped bands, confirming the presence of large amount of 3MMLCT character in the emissive state. Only the complexes bearing dfppy and dppa ligands reveal dual luminescence: a high‐energy structured emission originating from 3ILCT/3MLCT (intra‐ligand charge transfer/metal‐to‐ligand charge transfer) and an unstructured low‐energy band associated with 3MMLCT. In order to describe the nature of the electronic transitions, density functional theory calculations were performed for all the complexes.  相似文献   

19.
Heterometallic platinum complexes cover a huge field, as shown by a recent survey covering the crystallographic and structural data of almost 1500 examples. About 5% of those complexes exists as isomers and are summarized in this review; except one cis-trans example, the remainder are distortion isomers. These are discussed in terms of the coordination about the platinum atom, and correlations are drawn between donor atom, bond lengths and interbond angles, with attention to trans effect and metal-metal bonds. Distortion isomers, differing only by degree of distortion in Pt–L and Pt–M distances and L–M–L bond angles, spread over a wide range of oxidation states of platinum: zero, +1, +2 (most common) and +4. The mean Pt–Pt bond distance elongate with increase in oxidation state of platinum: 2.705 Å (Pt(0)–Pt(0)) < 2.720 Å (Pt(I)–Pt(I)) < 2.773 Å (Pt(II)–Pt(II)). The shortest mean Pt–M bond distances are: Pt(0)–Ga = 2.37 Å; Pt(I) = Au 2.697 Å, Pt(II)–Fe = 2.625 Å and Pt(IV)–Sn, 2.580 Å.  相似文献   

20.
The substitution reactions of the complexes [{trans‐Pt(NH3)2H2O}2(μ‐1,4‐diaminobutane)]4+ ( I ), [{trans‐Pt(NH3)2H2O}2(μ‐1,6‐diaminohexane)]4+ ( II ), and [{trans‐Pt(NH3)2H2O}2(μ‐1,8‐diaminooctane)]4+ ( III ), with nucleophiles L‐cysteine (L‐Cys), glutathione (GSH), guanosine‐5′‐monophosphate (5′‐GMP), L‐histidine (L‐His), and pyridine were studied in 0.1 M NaClO4 aqueous solutions at pH = 2.5. The substitutions were studied under pseudo‐first‐order conditions as a function of concentration and temperature using UV–vis spectrophotometry. At three different temperatures (288, 298, and 308 K) the reactions of the II and III complexes and 5′‐GMP were studied. The order of reactivity of study ligands is L‐Cys > GSH > 5′‐GMP > L‐His > pyridine and the order of reactivity of the complexes is I < II ≈ III . The obtained results indicate that the structure of the alkanediamine linker in the dinuclear Pt(II) complexes controls the substitution process. The negative values reported for entropy of activation confirmed the associative substitution mode. These results are discussed in order to find the connection between structure and reactivity of the dinuclear Pt(II) complexes.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号